Elasticity (physics)

Source: Wikipedia, the free encyclopedia.

In physics and materials science, elasticity is the ability of a body to resist a distorting influence and to return to its original size and shape when that influence or force is removed. Solid objects will deform when adequate loads are applied to them; if the material is elastic, the object will return to its initial shape and size after removal. This is in contrast to plasticity, in which the object fails to do so and instead remains in its deformed state.

The physical reasons for elastic behavior can be quite different for different materials. In metals, the atomic lattice changes size and shape when forces are applied (energy is added to the system). When forces are removed, the lattice goes back to the original lower energy state. For rubbers and other polymers, elasticity is caused by the stretching of polymer chains when forces are applied.

ideal concept
only; most materials which possess elasticity in practice remain purely elastic only up to very small deformations, after which plastic (permanent) deformation occurs.

In

yield strength is the maximum stress
that can arise before the onset of plastic deformation. Its SI unit is also the pascal (Pa).

Overview

When an elastic material is deformed due to an external force, it experiences internal resistance to the deformation and restores it to its original state if the external force is no longer applied. There are various elastic moduli, such as Young's modulus, the shear modulus, and the bulk modulus, all of which are measures of the inherent elastic properties of a material as a resistance to deformation under an applied load. The various moduli apply to different kinds of deformation. For instance, Young's modulus applies to extension/compression of a body, whereas the shear modulus applies to its shear.[1] Young's modulus and shear modulus are only for solids, whereas the bulk modulus is for solids, liquids, and gases.

The elasticity of materials is described by a

isotropic, the linearized stress–strain relationship is called Hooke's law, which is often presumed to apply up to the elastic limit for most metals or crystalline materials whereas nonlinear elasticity is generally required to model large deformations of rubbery materials even in the elastic range. For even higher stresses, materials exhibit plastic behavior, that is, they deform irreversibly and do not return to their original shape after stress is no longer applied.[3] For rubber-like materials such as elastomers, the slope of the stress–strain curve increases with stress, meaning that rubbers progressively become more difficult to stretch, while for most metals, the gradient decreases at very high stresses, meaning that they progressively become easier to stretch.[4] Elasticity is not exhibited only by solids; non-Newtonian fluids, such as viscoelastic fluids, will also exhibit elasticity in certain conditions quantified by the Deborah number. In response to a small, rapidly applied and removed strain, these fluids may deform and then return to their original shape. Under larger strains, or strains applied for longer periods of time, these fluids may start to flow like a viscous
liquid.

Because the elasticity of a material is described in terms of a stress–strain relation, it is essential that the terms stress and strain be defined without ambiguity. Typically, two types of relation are considered. The first type deals with materials that are elastic only for small strains. The second deals with materials that are not limited to small strains. Clearly, the second type of relation is more general in the sense that it must include the first type as a special case.

For small strains, the measure of stress that is used is the

isotropic media) is called the generalized Hooke's law. Cauchy elastic materials and hypoelastic materials are models that extend Hooke's law to allow for the possibility of large rotations, large distortions, and intrinsic or induced anisotropy
.

For more general situations, any of a number of

stress measures can be used, and it is generally desired (but not required) that the elastic stress–strain relation be phrased in terms of a finite strain measure that is work conjugate to the selected stress measure, i.e., the time integral of the inner product of the stress measure with the rate of the strain measure should be equal to the change in internal energy for any adiabatic process
that remains below the elastic limit.

Units

International System

The SI unit for elasticity and the elastic modulus is the pascal (Pa). This unit is defined as force per unit area, generally a measurement of pressure, which in mechanics corresponds to stress. The pascal and therefore elasticity have the dimension L−1⋅M⋅T−2.

For most commonly used engineering materials, the elastic modulus is on the scale of gigapascals (GPa, 109 Pa).

Linear elasticity

As noted above, for small deformations, most elastic materials such as

displacement
,

where k is a constant known as the rate or spring constant. It can also be stated as a relationship between

stress
and
strain
:

where E is known as the Young's modulus.[7]

Although the general proportionality constant between stress and strain in three dimensions is a 4th-order tensor called stiffness, systems that exhibit symmetry, such as a one-dimensional rod, can often be reduced to applications of Hooke's law.

Finite elasticity

The elastic behavior of objects that undergo finite deformations has been described using a number of models, such as

deformation gradient (F) is the primary deformation measure used in finite strain theory
.

Cauchy elastic materials

A material is said to be Cauchy-elastic if the

deformation gradient
F alone:

It is generally incorrect to state that Cauchy stress is a function of merely a

strain tensor
, as such a model lacks crucial information about material rotation needed to produce correct results for an anisotropic medium subjected to vertical extension in comparison to the same extension applied horizontally and then subjected to a 90-degree rotation; both these deformations have the same spatial strain tensors yet must produce different values of the Cauchy stress tensor.

Even though the stress in a Cauchy-elastic material depends only on the state of deformation, the work done by stresses might depend on the path of deformation. Therefore, Cauchy elasticity includes non-conservative "non-hyperelastic" models (in which work of deformation is path dependent) as well as conservative "hyperelastic material" models (for which stress can be derived from a scalar "elastic potential" function).

Hypoelastic materials

A hypoelastic material can be rigorously defined as one that is modeled using a constitutive equation satisfying the following two criteria:[8]

  1. The Cauchy stress at time depends only on the order in which the body has occupied its past configurations, but not on the time rate at which these past configurations were traversed. As a special case, this criterion includes a Cauchy elastic material, for which the current stress depends only on the current configuration rather than the history of past configurations.
  2. There is a tensor-valued function such that in which is the material rate of the Cauchy stress tensor, and is the spatial
    velocity gradient
    tensor.

If only these two original criteria are used to define hypoelasticity, then

deformation gradient
but do not start and end at the same internal energy.

Note that the second criterion requires only that the function exists. As detailed in the main hypoelastic material article, specific formulations of hypoelastic models typically employ so-called objective rates so that the function exists only implicitly and is typically needed explicitly only for numerical stress updates performed via direct integration of the actual (not objective) stress rate.

Hyperelastic materials

Hyperelastic materials (also called Green elastic materials) are conservative models that are derived from a

deformation gradient
via a relationship of the form

This formulation takes the energy potential (W) as a function of the

deformation gradient
(). By also requiring satisfaction of
Cauchy-Green deformation tensor
(), in which case the hyperelastic model may be written alternatively as

Applications

Linear elasticity is used widely in the design and analysis of structures such as

beams, plates and shells, and sandwich composites. This theory is also the basis of much of fracture mechanics
.

Hyperelasticity is primarily used to determine the response of elastomer-based objects such as gaskets and of biological materials such as soft tissues and cell membranes.

Factors affecting elasticity

In a given

isotropic solid, with known theoretical elasticity for the bulk material in terms of Young's modulus,the effective elasticity will be governed by porosity. Generally a more porous material will exhibit lower stiffness. More specifically, the fraction of pores, their distribution at different sizes and the nature of the fluid with which they are filled give rise to different elastic behaviours in solids.[9]

For

vibrations of the molecules, all of which are dependent on temperature.[12]

See also

Notes

  1. ^ Descriptions of material behavior should be independent of the geometry and shape of the object made of the material under consideration. The original version of Hooke's law involves a stiffness constant that depends on the initial size and shape of the object. The stiffness constant is therefore not strictly a material property.[citation needed]

References

  1. ^ Landau LD, Lipshitz EM. Theory of Elasticity, 3rd Edition, 1970: 1–172.
  2. .
  3. .
  4. .
  5. .
  6. ^ "Strength and Design". Centuries of Civil Engineering: A Rare Book Exhibition Celebrating the Heritage of Civil Engineering. Linda Hall Library of Science, Engineering & Technology. Archived from the original on 13 November 2010.[page needed]
  7. .
  8. .
  9. ^ Modeling of effective elastic properties International Journal of Solids and Structures Volume 162, 1 May 2019, Pages 36-44
  10. .
  11. .
  12. .

External links